Practical Synthesis of β-Acyl and β-Alkoxycarbonyl Heterocyclic

Mar 2, 2011 - Julia–Kocienski Reaction-Based 1,3-Diene Synthesis: Aldehyde-Dependent (E,E/E,Z)-Selectivity. François Billard , Raphaël Robiette , ...
0 downloads 0 Views 986KB Size
NOTE pubs.acs.org/joc

Practical Synthesis of β-Acyl and β-Alkoxycarbonyl Heterocyclic Sulfones Jirí Pospísil* and Hitoshi Sato† Institute of Condensed Matter and Nanosciences-Molecules, Solids and Reactivity (IMCN/MOST), Universite catholique de Louvain, B^atiment Lavoisier, Place Louis Pasteur 1, B-1348 Louvain-la-Neuve, Belgium

bS Supporting Information ABSTRACT: A short and efficient synthesis for β-acyl and βalkoxycarbonyl heterocyclic sulfones containing benzothiazol (BT) and phenyltetrazol (PT) heterocyclic core is presented here. The method seems to be general and provides the desired C-nucleophiles in very good to excellent yields from readily available starting materials.

O

ver the past two decades, asymmetric organocatalysis has developed into a field permitting the elegant introduction of chiral information into a plethora of various substrates.1 During this period, several efficient organocatalytic protocols that create a novel C-C bond have been developed. The asymmetric organocatalytic conjugate addition of carbon nucleophiles to R,β-unsaturated carbonyl substrates is one typical example.2 It is therefore surprising that efficient protocols allowing the addition of alkynylic3 and alkenylic3,4 functionalities have only been recently developed by MacMillan4 and Jørgensen.3 Even though both of these protocols provide the desired alkenylic derivatives in excellent yields and enantioselectivities, they are conceptually different. MacMillan’s approach is based on the use of vinyl-type borate complexes5 that acts as nucleophiles during the organocatalytic asymmetric conjugate addition process. The Jørgensen approach, on the other hand, is based on a two-step tandem process (see Scheme 1). First, an organocatalyst-promoted conjugate addition of β-keto heterocyclic sulfone 2 nucleophile to R,β-unsaturated carbonyl compound generates adduct 3, which is then transformed to olefin 4 (via reduction/Smile rearrangement6 protocol) or alkyne 5 (via enolization/Smile rearrangement protocol6). Later on, Jørgensen et al. used the same β-keto heterocyclic sulfones 2 to perform a formal transition-metal-free Sonogashira coupling and R-carbonyl arylation reaction7 (Scheme 1). Finally, if β-acyl-heterocyclic sulfones 9 are used as reagents in the Julia-Kocienski olefination reaction, R,β-unsaturated esters 10 can be prepared (Scheme 2).8 In our laboratory we recently started several synthetic ventures that use and/or are based on the compounds that contains β-keto and/or β-alkoxycarbonyl motifs, and to our great surprise, we found out that in the literature there is only one synthetic approach that yields sulfones 2 or 9. In this method, sulfones 2 and 9 are prepared in two steps starting from R-bromo (halo) ketones or esters as the key intermediates toward 2 or 9 (Scheme 3). Unfortunately, in our case the synthesis of the desired R-halo carbonyl compounds proved r 2011 American Chemical Society

to be long and tedious, and therefore we decided to evaluate a new synthetic approach to β-carbonyl heterocyclic sulfones that would better fit our purposes. To achieve the most versatile approach to sulfones of general structure 11, we decided to base the synthesis on the pairing of sulfone 13 with electrophile 14 (Scheme 4). Even though this type of disconnection seems to be obvious and is easily applicable to β-keto phenyl sulfone synthesis,9 this approach is not easily applicable if β-keto BT- and PT-sulfones are wanted. The reason is that both BT- and PT-sulfones contain one electrophilic center within their heterocyclic core. As a consequence, if BT- or PT-sulfones 13 are treated with a non-nucleophilic base,10 selfcondensation occurs (Scheme 5).11 Thus if we want to prepare the desired β-carbonyl sulfones 11 starting from sulfone 13 and carbonyl electrophile 14, conditions under which nucleophiles present in the reaction mixture are unreactive toward the nucleophilic center have to be found. To find suitable coupling conditions, we decided to study a pairing between sulfone 13a and benzoyl chloride 14a (Table 1). Initially, the Li anion generated from 13a was prepared with the use of the non-nucleophilic base (LiN(TMS)2) at low temperature, and BzCl, an electrophile, was added after 30 min (Table 1, entry 1). In this case, the desired product 11a was obtained only in very low yield. Since the control experiment showed that the low yield is not caused by instability or self-condensation of the intermediate16,12 it remains that the low yield is caused by selfcondensation of 13a. To avoid this side reaction, the mixing time of 13a with the base was progressively diminished (Table 1, entries 2 and 3). Gratifyingly, it was observed that the desired product 11a was isolated in excellent yield (96%) if generated sulfone anion 13a was quenched rapidly (BzCl was added immediately after the base).13

Received: November 23, 2010 Published: March 02, 2011 2269

dx.doi.org/10.1021/jo102326p | J. Org. Chem. 2011, 76, 2269–2272

The Journal of Organic Chemistry Scheme 1. Application of β-Keto-heterocyclic Sulfone 2 in Organic Synthesis

NOTE

Scheme 5. Alkyl BT-Sulfone Base-Mediated SelfCondensation

Table 1. Optimization of the Reaction Conditions

Scheme 2. β-Acyl-heterocyclic Sulfone 9 in the Context of a Julia-Kocienski Reaction

Scheme 3. Standard Approach to β-Carbonyl Heterocyclic Sulfones 11

Scheme 4. Retrosynthetic Approach to β-Carbonyl Heterocyclic Sulfones 11

a

At this stage we decided to test the coupling reaction with other types of benzoyl electrophiles to evaluate its scope and limitations. Unsurprisingly benzoylating agents with reactivity similar to that of benzoyl chloride, such as BzF, BzBr, Bz2O, or benzoyl imidazole14 (Table 1, entries 7, 9-11), yielded the desired βketo sulfone 11a in essentially the same yield. On the other hand, methyl benzoyl ester proved to be unreactive under our reaction

conditions (Table 1, entry 8). In this case, only the self-condensation product of 13a (compound 15) was observed.15 Having devised suitable reaction conditions to prepare this β-acyl sulfone, its scope and limitations were explored. A selection of pertinent results is shown in Table 2. Both alkyl and aryl acyl derivatives react smoothly with lithiated sulfones 13 (Table 2). In all cases the reaction yields were more than 80% except when sulfone 13b was reacted with monoethyl oxalyl chloride (Table 2, entry 8). In this case the desired product 11f was obtained in 78% yield. Even acyl chlorides, bearing enolizable hydrogen atoms in the R position, reacted under the given reaction conditions to yield the desired R-acyl sulfones in very good yields. 2270

Overall yields refer to pure, isolated products.

dx.doi.org/10.1021/jo102326p |J. Org. Chem. 2011, 76, 2269–2272

The Journal of Organic Chemistry Table 2. Preparation of Acyl Heterocyclic Sulfones 11 Starting from Sulfones 13 and Acyl-Containing Electrophiles

At this stage this protocol was extended to the synthesis of β-alkoxy carbonyl sulfones 17 (Table 3). For this purpose, three different types of alkoxy carbonylating reagents (bearing Cl, imidazole,16 or OCOR as a leaving group) were tested. The nature of the leaving group was shown to have little effect on the reaction yields, and all three electrophiles might be used as coupling partners. Additionally, the reaction conditions tolerate various functionalities, e.g., TBDPS ethers, phenyl ethers, and halogenated or unsaturated alkanes (Table 2, entries 7-10 and Table 3, entries 12-18). In summary, we have uncovered a short and efficient approach to R-acyl and R-alkoxy carbonyl heterocyclic sulfones 11 and 17, respectively, starting from heterocyclic sulfones and acyl or alkoxy carbonyl derivatives. We believe that this general

NOTE

Table 3. Preparation of Alkoxycarbonyl Heterocyclic Sulfones 11 Starting from Sulfones 13 and Alkoxy CarbonylContaining Electrophiles

approach to making this class of C-nucleophiles, which can be easily transformed into olefins or alkynes, will extend their use beyond the field of asymmetric organocatalysis. Further development and use of β-carbonyl heterocyclic sulfones of general structure 11 and 17 is now in progress in our laboratory.

’ EXPERIMENTAL SECTION Representative Procedure for the Preparation of 2-(Benzo[d]thiazol-2-ylsulfonyl)-1-phenylethanone (11a). A solution of 2271

dx.doi.org/10.1021/jo102326p |J. Org. Chem. 2011, 76, 2269–2272

The Journal of Organic Chemistry sulfone 13a (100 mg, 0.47 mmol, 1.0 equiv) in THF (2.4 mL, 0.20M) was cooled to -78 °C, and LiN(TMS)2 (1.0 M solution in THF) (1.03 mL, 1.03 mmol, 2.2 equiv) was added dropwise. The color of the reaction mixture turned from colorless or slightly yellow to orange within approximately 10-20 s. Immediately after, a solution of benzoyl chloride (60 μL, 0.52 mmol, 1.1 equiv) in THF (0.25 mL) was added. The color of the reaction mixture faded within 1 to 5 min after the benzoyl chloride addition. The resulting mixture was stirred at -78 °C for 30 min, allowed to warm to 0 °C within 1 h, and stirred at 0 °C for a further 30 min before a saturated aqueous solution of NH4Cl (7.5 mL) was added. The whole mixture was extracted with EtOAc (3  40 mL); the combined organic layers were washed with brine (25 mL), dried over MgSO4, and filtered; and the solvents were removed under reduced pressure. The residue was purified by flash column chromatography on SiO2 (petroleum ether/EtOAc = 4:1 f 2:1 f 1:1), and the reaction yielded 143 mg (96%) of 11a as slightly yellow solid: mp 123-124 °C; 1H NMR (300 MHz, CDCl3) δ 5.22 (s, 2H), 7.43-7.53 (m, 2H), 7.55-7.69 (m, 3H), 7.94 (dd, J = 8.4 Hz, J = 1.2 Hz, 2H), 8.01 (dd, J = 7.0 Hz, J = 2.2 Hz, 1H), 8.20 (dd, J = 7.2 Hz, J = 2.1 Hz, 1H); 13C NMR (75 MHz, CDCl3) δ 61.4, 122.6, 125.7, 127.9, 128.4, 129.17, 129.19, 134.9, 135.6, 137.3, 152.6, 165.5, 187.3; IR (neat) ν-1 1683 (s); MS (APCI) (relative intensity) m/z 318 (Mþ þ 1, 100), 319 (20), 236 (9), 105 (11). Anal. Calcd for C15H11NO3S2: C, 56.76; H, 3.49; N, 4.41. Found: C, 56.78; H, 3.11; N, 4.67.

’ ASSOCIATED CONTENT

bS

Supporting Information. Spectroscopic and analytical data for all new compounds, as well as experimental procedures. This material is available free of charge via the Internet at http://pubs.acs.org.

NOTE

(3) (a) Paix~ao, M. W.; Holub, N.; Vila, C.; Nielsen, M.; Jørgensen, K. A. Angew. Chem., Int. Ed. 2009, 48, 7338. (b) Nielsen, M.; Jacobsen, C. B.; Paix~ao, M. W.; Holub, N.; Jørgensen, K. A. J. Am. Chem. Soc. 2009, 131, 10581. (4) Lee, S.; MacMillan, D. W. C. J. Am. Chem. Soc. 2007, 129, 15438. (5) (a) Molander, G. A.; Fumagalli, T. J. Org. Chem. 2006, 71, 5743. (b) Vedejs, E.; Chapman, R. W.; Fields, S. C.; Lin, S.; Schrimpf, M. R. J. Org. Chem. 1995, 60, 3020. (6) Warren, L. A.; Smiles, S. J. Chem. Soc. 1930, 1327. (7) Pr€uger, B.; Hofmeister, G. E.; Jacobsen, C. B.; Alberg, D. G.; Nielsen, M.; Jørgensen, K. A. Chem.—Eur. J. 2010, 16, 3783. (8) (a) Blakemore, P. R.; Ho, D. K. H.; Mieke Nap, W. Org. Biomol. Chem. 2005, 3m, 1365. (b) Giesbrecht, H. E.; Knight, B. J.; Tanguileg., N. R.; Emerson, C. R.; Blakemore, P. R. Synlett 2010, 374. (9) For application of this approach for the synthesis β-carbonyl phenyl sulfones, see: Bartlett, B. A.; Green, F. R., III; Rose, E. H. J. Am. Chem. Soc. 1978, 100, 4852. (10) If nucleophilic bases, e.g., n-BuLi, are used, the competition between its addition to heterocyclic core of the sulfone and deprotonation of the hydrogen atom R to sulfone is observed. (11) Baudin, J. B.; Hareau, G.; Julia, S. A.; Lorne, R.; Ruel, O. Bull. Soc. Chim. Fr. 1993, 130, 856. (12) See Supporting Information. (13) For discussion about Barbier-type reaction conditions, see Supporting Information. (14) Prepared in situ from the corresponding carboxylic acid: Ibarra, C. A.; Rodriguez, R. C.; Fernandez Monreal, M. C.; Garcia Navarro, F. J.; Martin Tesorero, J. J. Org. Chem. 1989, 54, 5620. (15) Coupling between sulfone 13a and BzOMe was extensively studied using various reaction conditions, bases, and solvents. In all cases, only traces of the desired product 11a were observed. (16) (a) Rannard, S. P.; Davis, N. J. Org. Lett. 1999, 1, 933. (b) Rannard, S. P.; Davis, N. J. Org. Lett. 2000, 2, 2117.

’ AUTHOR INFORMATION Corresponding Author

*E-mail: [email protected]. Notes †

Visiting researcher from Tohoku University, Japan.

’ ACKNOWLEDGMENT J.P. is grateful to Prof. Istvan E. Marko (U.C.L.) for his continuous support. We gratefully acknowledge the Universite catholique de Louvain, the Fond de la Recherche Scientific (Charge de Recherche F.S.R.-FNRS to J.P.), Socrates-Erasmus exchange program (H.S.), and Tohoku University G-COE program (IREMC) (H.S.) for financial support. ’ REFERENCES (1) (a) Nielsen, M.; Jacobsen, C. B.; Holub, N.; Paixxao, M. W.; Jørgensen, K. A. Angew. Chem., Int. Ed. 2010, 49, 2668. (b) Melchiorre, P.; Marigo, M.; Carlone, A.; Bartolli, G. Angew. Chem., Int. Ed. 2008, 47, 6138. (c) Dondoni, A.; Massi, A. Angew. Chem., Int. Ed. 2008, 47, 4638.(d) Chem. Rev. 2007, 107 (12), special issue on organocatalysis. (2) For recent examples, see: (a) Luo, J.; Xu, L.-W.; Hay, R. A. S.; Lu, Y. Org. Lett. 2009, 11, 437. (b) Prakash, G. K. S.; Wang, F.; Steward, T.; Mathew, T.; Olah, G. A. Proc. Natl. Acad. Sci. U.S.A. 2009, 106, 4090. (c) Cassani, C.; Bernardi, L.; Fini, F.; Ricci, A. Angew. Chem., Int. Ed. 2009, 48, 5694. (d) Cid, M. B.; Cantarero, J. L.; Duce, S.; Ruano, J. L. G. J. Org. Chem. 2009, 74, 431. (e) Furukawa, T.; Shibata, N.; izuta, S.; Nakamura, S.; Toru, T.; Shiro, M. Angew. Chem., Int. Ed. 2008, 47, 8051. (f) Huang, H.; Jacobsen, E. N. J. Am. Chem. Soc. 2006, 128, 7170. (g) Pulkkinnen, J.; Aburel, P. S.; Halland, N.; Jørgensen, K. A. Adv. Synth. Catal. 2004, 346, 1077. 2272

dx.doi.org/10.1021/jo102326p |J. Org. Chem. 2011, 76, 2269–2272